• Aucun résultat trouvé

Limit theorems for nonlinear functionals of Volterra processes via white noise analysis

N/A
N/A
Protected

Academic year: 2021

Partager "Limit theorems for nonlinear functionals of Volterra processes via white noise analysis"

Copied!
32
0
0

Texte intégral

(1)

DOI:10.3150/10-BEJ258

Limit theorems for nonlinear functionals of Volterra processes via white noise analysis

S É BA S T I E N DA R S E S1, I VA N N O U R D I N2and DAV I D N UA L A RT3

1Université Aix-Marseille I, 39 rue Joliot Curie, 13453 Marseille Cedex 13, France.

E-mail:darses@cmi.univ-mrs.fr

2Laboratoire de Probabilités et Modèles Aléatoires, Université Pierre et Marie Curie (Paris VI), Boîte Courrier 188, 4 place Jussieu, 75252 Paris Cedex 05, France. E-mail:ivan.nourdin@upmc.fr

3Department of Mathematics, University of Kansas, Lawrence, KS 66045, USA.

E-mail:nualart@math.ku.edu

By means of white noise analysis, we prove some limit theorems for nonlinear functionals of a given Volterra process. In particular, our results apply to fractional Brownian motion (fBm) and should be com- pared with the classical convergence results of the 1980s due to Breuer, Dobrushin, Giraitis, Major, Surgailis and Taqqu, as well as the recent advances concerning the construction of a Lévy area for fBm due to Coutin, Qian and Unterberger.

Keywords:fractional Brownian motion; limit theorems; Volterra processes; white noise analysis

1. Introduction

FixT >0 and letB=(Bt)t0be a fractional Brownian motion with Hurst indexH(0,1), defined on some probability space(,B, P ). Assume thatBis the completedσ-field generated byB. Fix an integerk≥2 and, forε >0, consider

Gε=εk(1H ) T

0

hk

Bu+εBu εH

du. (1.1)

Here, and in the rest of this paper,

hk(x)=(−1)kex2/2 dk

dxk(ex2/2) (1.2)

stands for thekth Hermite polynomial. We haveh2(x)=x2−1,h3(x)=x3−3xand so on.

Since the seminal works [3,6,7,19,20] by Breuer, Dobrushin, Giraitis, Major, Surgailis and Taqqu, the following three convergence results are classical:

• ifH <1−2k1, then

(Bt)t∈[0,T], εk(1H )1/2GεLaw

−→ε0

(Bt)t∈[0,T], N

, (1.3)

whereN∼N(0, T ×k!T

0 ρk(x)dx)is independent ofB, withρ(x)=21(|x+1|2H+ |x− 1|2H−2|x|2H);

1350-7265 © 2010 ISI/BS

(2)

• ifH=1−2k1, then

(Bt)t∈[0,T], Gε log(1/ε)

Law

−→ε0

(Bt)t∈[0,T], N

, (1.4)

whereN∼N(0, T ×2k!(12k1)k(11k)k)is independent ofB;

• ifH >1−2k1, then

GεL

2()

−→ε0Z(k)T , (1.5)

whereZT(k)denotes the Hermite random variable of orderk; see Section4.1for its definition.

Combining (1.3) with the fact that sup0<ε1E[|εk(1H )1/2Gε|p]<∞for allp≥1 (use the boundedness of Var(εk(1H )1/2Gε)and a classical hypercontractivity argument), we have, for allηL2()and ifH <1−2k1, that

εk(1H )1/2E[ηGε] −→

ε0E(ηN )=E(η)E(N )=0

(a similar statement holds in the critical case H =1− 2k1). This means that εk(1H )1/2Gε convergesweaklyinL2()to zero. The following question then arises. Is there a normalization ofGεensuring that it converges weaklytowards anon-zerolimit whenH≤1−2k1? If so, then what can be said about the limit? The first purpose of the present paper is to provide an answer to this question in the framework ofwhite noise analysis.

In [14], it is shown that for allH(0,1), the time derivativeB˙ (called thefractional white noise) is a distribution in the sense of Hida. We also refer to Bender [1], Biaginiet al.[2] and references therein for further works on the fractional white noise.

Since we haveE(Bu+εBu)2=ε2H, observe thatGεdefined in (1.1) can be rewritten as Gε=

T 0

Bu+εBu ε

k

du, (1.6)

where(. . .)k denotes thekth Wick product. In Proposition9 below, we will show that for all H(121k,1),

εlim0

T

0

Bu+εBu ε

k

du= T

0

B˙ukdu, (1.7)

where the limit is in the(S)sense.

In particular, we observe two different types of asymptotic results forGε when H(12

1

k,1−2k1): convergence (1.7) in(S)to a Hida distribution, and convergence (1.3) in law to a normal law, with rateε1/2k(1H ). On the other hand, whenH(12k1,1), we obtain from (1.5) that the Hida distributionT

0 B˙skdsturns out to be the square-integrable random variable ZT(k), which is an interesting result in its own right.

(3)

In Proposition 4, the convergence (1.7) in (S) is proved for a general class of Volterra processes of the form

t 0

K(t, s)dWs, t≥0, (1.8)

whereW stands for a standard Brownian motion, provided the kernelKsatisfies some suitable conditions; see Section3.

We also provide a new proof of the convergence (1.3) based on the recent general criterion for the convergence in distribution to a normal law of a sequence of multiple stochastic integrals established by Nualart and Peccati [15] and by Peccati and Tudor [17], which avoids the classical method of moments.

In two recent papers [9,10], Marcus and Rosen have obtained central and non-central limit theorems for a functional of the form (1.1), where B is a mean zero Gaussian process with stationary increments such that the covariance function ofB, defined byσ2(|ts|)=Var(BtBs), is either convex (plus some additional regularity conditions), concave or given byσ2(h)= hr with 1< r <2. These theorems include the convergence (1.3) and, unlike our simple proof, are based on the method of moments.

In the second part of the paper, we develop a similar analysis for functionals of two indepen- dent fractional Brownian motions (or, more generally, Volterra processes) related to the Lévy area. More precisely, consider twoindependentfractional Brownian motionsB(1)andB(2)with (for simplicity) the same Hurst indexH(0,1). We are interested in the convergence, asε→0, of

Gε:=

T 0

Bu(1)Bu(2)+εBu(2)

ε du (1.9)

and

G˘ε:=

T

0

u 0

Bv(1)+εBv(1)

ε dv

Bu(2)+εBu(2)

ε du. (1.10)

Note that Gε coincides with the ε-integral associated with the forward Russo–Vallois integral T

0 B(1)dB(2); see, for example, [18] and references therein. Over the last decade, the conver- gences ofGε andG˘ε (or of related families of random variables) have been intensively studied.

Sinceε1u

0(Bv(1)+εBv(1))dv converges pointwise toBu(1) for anyu, we could think that the asymptotic behaviors ofGε andG˘ε are very close asε→0. Surprisingly, this is not the case.

Actually, only the result forG˘εagrees with the seminal result of Coutin and Qian [4] (that is, we have convergence ofG˘ε inL2()if and only ifH >1/4) and with the recent result by Unter- berger [21] (that is, adequately renormalized,G˘εconverges in law ifH <1/4). More precisely:

• ifH <1/4, then

Bt(1), Bt(2)

t∈[0,T], ε1/22HG˘εLaw

−→ε0

Bt(1), Bt(2)

t∈[0,T], N

, (1.11)

(4)

whereN∼N(0, Tσ˘H2)is independent of(B(1), B(2))and

˘

σH2 = 1

4(2H+1)(2H+2)

R(|x+1|2H+ |x−1|2H−2|x|2H)

×(2|x|2H+2− |x+1|2H+2− |x−1|2H+2)dx;

• ifH=1/4, then

Bt(1), Bt(2)

t∈[0,T], G˘ε log(1/ε)

Law

−→ε0

Bt(1), Bt(2)

t∈[0,T], N

, (1.12)

whereN∼N(0, T /8)is independent of(B(1), B(2));

• ifH >1/4, then

G˘εL

2()

−→ε0

T 0

Bu(1) ˙Bu(2)du= T

0

Bu(1)dBu(2); (1.13)

• for allH(0,1), we have

G˘ε−→(S)

ε0

T 0

Bu(1) ˙Bu(2)du. (1.14)

However, forGε, we have, in contrast:

• ifH <1/2, then

Bt(1), Bt(2)

t∈[0,T], ε1/2HGεLaw

−→ε0

Bt(1), Bt(2)

t∈[0,T], N×S

, (1.15)

where

S=

0

(|x+1|2H+ |x−1|2H−2|x|2H)dx× T

0

Bu(1)2

du andN∼N(0,1), independent of(B(1), B(2));

• ifH≥1/2, then

GεL

2()

−→ε0

T

0

Bu(1) ˙Bu(2)du= T

0

Bu(1)dBu(2); (1.16)

• for allH(0,1), we have

Gε−→(S)

ε0

T

0

Bu(1) ˙Bu(2)du. (1.17)

Finally, we study the convergence, asε→0, of the so-called ε-covariation(following the terminology of [18]) defined by

Gε:=

T

0

Bu(1)+εBu(1)

ε ×Bu(2)+εBu(2)

ε du (1.18)

(5)

and we get:

• ifH <3/4, then

Bt(1), Bt(2)

t∈[0,T], ε3/22HGε

Law

−→ε0

Bt(1), Bt(2)

t∈[0,T], N

(1.19) withN∼N(0, TσH2)independent of(B(1), B(2))and

σH2 =1

4

R(|x+1|2H+ |x−1|2H−2|x|2H)2dx;

• ifH=3/4, then

Bt(1), Bt(2)

t∈[0,T], Gε

log(1/ε) Law

−→ε0

Bt(1), Bt(2)

t∈[0,T], N

(1.20) withN∼N(0,9T /32)independent of(B(1), B(2));

• ifH >3/4, then

GεL

2()

−→ε0

T

0

B˙u(1) ˙Bu(2)du; (1.21)

• for allH(0,1), we have

Gε−→(S)

ε0

T 0

B˙u(1) ˙Bu(2)du. (1.22) The paper is organized as follows. In Section2, we introduce some preliminaries on white noise analysis. Section3 is devoted to the study, using the language and tools of the previous section, of the asymptotic behaviors ofGε,GεandGεin the (more general) context whereBis a Volterra process. Section4is concerned with the fractional Brownian motion case. In Section5 (resp., Section6), we prove (1.3) and (1.4) (resp., (1.11), (1.12), (1.15), (1.19) and (1.20)).

2. White noise functionals

In this section, we present some preliminaries on white noise analysis. The classical approach to white noise distribution theory is to endow the space of tempered distributionsS(R)with a Gaussian measurePsuch that, for any rapidly decreasing functionηS(R),

S(R)

eix,ηP(dx)=e−|η|20/2.

Here,| · |0denotes the norm inL2(R)and·,·the dual pairing betweenS(R)andS(R). The existence of such a measure is ensured by Minlos’ theorem [8].

In this way, we can consider the probability space (,B,P), where=S(R). The pair- ing x, ξ can be extended, using the norm of L2(), to any function ξL2(R). Then,

(6)

Wt= ·,1[0,t]is a two-sided Brownian motion (with the convention that1[0,t]= −1[t,0]ift <0) and for anyξL2(R),

·, ξ =

−∞ξdW=I1(ξ ) is the Wiener integral ofξ.

LetL2(). The classical Wiener chaos expansion ofsays that there exists a sequence of symmetric square-integrable functionsφnL2(Rn)such that

= n=0

Inn), (2.1)

whereIndenotes the multiple stochastic integral.

2.1. The space of Hida distributions

Let us recall some basic facts concerning tempered distributions. Letn)n=0be the orthonormal basis ofL2(R)formed by the Hermite functions given by

ξn(x)1/4(2nn!)1/2ex2/2hn(x), x∈R, (2.2) wherehnare the Hermite polynomials defined in (1.2). The following two facts can immediately be checked: (a) there exists a constantK1>0 such that ξnK1(n+1)1/12; (b) since ξn=

n

2ξn1

n+1

2 ξn+1, there exists a constantK2>0 such thatξnK2n5/12.

Consider the positive self-adjoint operatorA (whose inverse is Hilbert–Schmidt) given by A= −dxd22 +(1+x2). We haveAξn=(2n+2)ξn.

For anyp≥0, define the spaceSp(R)to be the domain of the closure ofAp. Endowed with the norm |ξ|p:= |Apξ|0, it is a Hilbert space. Note that the norm| · |p can be expressed as follows, if one uses the orthonormal basisn):

|ξ|2p= n=0

ξ, ξn2(2n+2)2p.

We denote bySp(R)the dual ofSp(R). The norm inSp(R)is given by (see [16], Lemma 1.2.8)

|ξ|2p= n=0

|ξ, Apξn|2= n=0

ξ, ξn2(2n+2)2p

for anyξSp(R). One can show that the projective limit of the spacesSp(R),p≥0, isS(R), that the inductive limit of the spacesSp(R),p≥0, isS(R)and that

S(R)L2(R)S(R)

(7)

is a Gel’fand triple.

We can now introduce the Gel’fand triple

(S)L2()(S),

via the second quantization operator(A). This is an unbounded and densely defined operator onL2()given by

(A)=

n=0

In(Anφn),

wherehas the Wiener chaos expansion (2.1). Ifp≥0, then we denote by(S)p the space of random variablesL2()with Wiener chaos expansion (2.1) such that

pp:=E[|(A)p|2] = n=0

n!|φn|2p<.

In the above formula,|φn|p denotes the norm inSp(R)n. The projective limit of the spaces (S)p,p≥0, is called the space of test functions and is denoted by(S). The inductive limit of the spaces(S)p,p≥0, is called the space of Hida distributions and is denoted by(S). The elements of(S) are calledHida distributions. The main example is the time derivative of the Brownian motion, defined asW˙t= ·, δt. One can show that|δt|p<∞for somep >0.

We denote by, the dual pairing associated with the spaces(S)and(S). On the other hand (see [16], Theorem 3.1.6), for any(S), there existφnS(Rn)such that

, =

n=0

n!φn, ψn,

where=

n=0Inn)(S). Moreover, there existsp >0 such that 2p=

n=0

n!|φn|2p.

Then, with a convenient abuse of notation, we say thathas a generalized Wiener chaos expan- sion of the form (2.1).

2.2. The S -transform

A useful tool to characterize elements in(S) is the S-transform. The Wick exponential of a Wiener integralI1(η),ηL2(R), is defined by

:eI1(η): =eI1(η)−|η|20/2.

(8)

TheS-transform of an element(S)is then defined by S()(ξ )=

,:eI1(ξ ): ,

whereξS(R). One can easily see that theS-transform is injective on(S).

IfL2(), thenS()(ξ )=E[:eI1(ξ ):]. For instance, theS-transform of the Wick ex- ponential is

S

:eI1(η):

(ξ )=eη,ξ. Also,S(Wt)(ξ )=t

0ξ(s)dsandS(W˙t)(ξ )=ξ(t ).

Suppose that(S)has a generalized Wiener chaos expansion of the form (2.1). Then, for anyξS(R),

S()(ξ )=

n=0

φn, ξn, where the series converges absolutely (see [16], Lemma 3.3.5).

The Wick product of two functionals=

n=0Inn)and=

n=0Inn)belonging to (S)is defined as

=

n,m=0

In+mnφm).

It can be proven that(S). The following is an important property of theS-transform:

S()(ξ )=S()(ξ )S()(ξ ). (2.3)

If,andbelong toL2(), then we haveE[] =E[]E[].

The following is a useful characterization theorem.

Theorem 1. A functionF is theS-transform of an element(S)if and only if the following conditions are satisfied:

(1) for anyξ, ηS,zF (zξ+η)is holomorphic onC;

(2) there exist non-negative numbersK, aandpsuch that for allξS,

|F (ξ )| ≤Kexp(a|ξ|2p).

Proof. See [8], Theorems 8.2 and 8.10.

In order to study the convergence of a sequence in(S), we can use itsS-transform, by virtue of the following theorem.

Theorem 2. Let n(S) and Sn=S(n).Then, n converges in (S) if and only if the following conditions are satisfied:

(1) limn→∞Sn(ξ )exists for eachξS;

(9)

(2) there exist non-negative numbersK, aandpsuch that for alln∈N,ξ(S),

|Sn(ξ )| ≤Kexp(a|ξ|2p).

Proof. See [8], Theorem 8.6.

3. Limit theorems for Volterra processes

3.1. One-dimensional case

Consider a Volterra processB=(Bt)t0of the form Bt=

t

0

K(t, s)dWs, (3.1)

whereK(t, s)satisfiest

0K(t, s)2ds <∞for allt >0 andW is the Brownian motion defined on the white noise probability space introduced in the last section. Note that theS-transform of the random variableBt is given by

S(Bt)(ξ )= t

0

K(t, s)ξ(s)ds (3.2)

for anyξS(R). We introduce the following assumptions on the kernelK:

(H1) Kis continuously differentiable on{0< s < t <∞}and for anyt >0, we have t

0

∂K

∂t (t, s)

(ts)ds <∞;

(H2) k(t )=t

0K(t, s)dsis continuously differentiable on(0,).

Consider the operatorK+defined by K+ξ(t )=k(t)ξ(t)+

t

0

∂K

∂t (t, r)

ξ(r)ξ(t ) dr,

wheret >0 andξS(R). From Theorem1, it follows that the linear mappingξK+ξ(t )is theS-transform of a Hida distribution. More precisely, according to [14], define the function

C(t )= |k(t)| + t

0

∂K

∂t (t, r)

(tr)dr, t≥0, (3.3)

(10)

and observe that the following estimates hold (recall the definition (2.2) ofξn):

|K+ξ(t )| ≤C(t )(ξ+ ξ)

C(t ) n=0

|ξ, ξn|n+ ξn)

C(t )M n=0

|ξ, ξn|(n+1)5/12 (3.4)

C(t )M

n=0

|ξ, ξn|2(2n+2)17/6

n=0

(n+1)2

=C(t )M|ξ|17/12

for some constantsM >0 whose values are not always the same from one line to the next.

We have the following preliminary result.

Lemma 3. Fix an integerk≥1.LetB be a Volterra process with kernelK satisfying the con- ditions(H1)and (H2).Assume,moreover,that C defined by(3.3)belongs to Lk([0, T]).The functionξT

0 (K+ξ(s))kds is then the S-transform of an element of(S).This element is denoted byT

0 B˙ukdu.

Proof. We use Theorem1. Condition (1) therein is immediately checked, while for condition (2), we just write, using (3.4),

T

0

(K+ξ(s))kds ≤

T

0

|K+ξ(s)|kds≤M|ξ|17/12

T

0

Ck(s)ds.

Fix an integerk≥1 and consider the following, additional, condition.

(Hk3) The maximal functionD(t)=sup0<εε0 1εt+ε

t C(s)ds belongs toLk([0, T])for any T >0 and for someε0>0.

We can now state the main result of this section.

Proposition 4. Fix an integerk≥1.LetB be a Volterra process with kernelK satisfying the conditions(H1),(H2)and(Hk3).The following convergence then holds:

T 0

Bu+εBu ε

k

du−→(S)

ε0

T 0

B˙ukdu.

(11)

Proof. FixξS(R)and set Sε(ξ )=S

T 0

Bu+εBu ε

k

du

(ξ ).

From linearity and property (2.3) of theS-transform, we obtain Sε(ξ )=

T

0

(S(Bu+εBu)(ξ ))k

εk du. (3.5)

Equation (3.2) yields

S(Bu+εBu)(ξ )= u+ε

0

K(u+ε, r)ξ(r)dr− u

0

K(u, r)ξ(r)dr. (3.6) We claim that

u+ε

0

K(u+ε, r)ξ(r)dr− u

0

K(u, r)ξ(r)dr= u+ε

u

K+ξ(s)ds. (3.7) Indeed, we can write

u+ε

u

K+ξ(s)ds= u+ε

u

k(s)ξ(s)ds+ u+ε

u

s

0

∂K

∂s (s, r)

ξ(r)ξ(s) dr

ds

(3.8)

=A(1)u +A(2)u . We have, using Fubini’s theorem, that

A(2)u = − u+ε

u

ds s

0

dr∂K

∂s(s, r) s

r

dθ ξ(θ )

(3.9)

= − u+ε

0

dθ ξ(θ ) θ

0

dr

K(u+ε, r)K(θu, r) .

This can be rewritten as A(2)u = −

u

0

K(u+ε, r)K(u, r)

ξ(u)ξ(r) dr

(3.10)

u+ε

u

dθ ξ(θ ) θ

0

dr

K(u+ε, r)K(θ, r) .

On the other hand, integration by parts yields A(1)u =ξ(u+ε)

u+ε

0

K(u+ε, r)dr

(3.11)

ξ(u) u

0

K(u, r)dr− u+ε

u

ds ξ(s) s

0

dr K(s, r).

(12)

Therefore, adding (3.11) and (3.10) yields A(1)u +A(2)u =ξ(u+ε)

u+ε

0

K(u+ε, r)dr−ξ(u) u

0

K(u, r)dr

u

0

K(u+ε, r)K(u, r)

ξ(u)ξ(r)

dr (3.12)

u+ε

u

dθ ξ(θ ) θ

0

K(u+ε, r)dr.

Note that, by integrating by parts, we have

u+ε

u

dθ ξ(θ ) θ

0

K(u+ε, r)dr

= −ξ(u+ε) u+ε

0

K(u+ε, r)dr+ξ(u) u

0

K(u+ε, r)dr (3.13) +

u+ε

u

K(u+ε, r)ξ(r)dr.

Thus, substituting (3.13) into (3.12), we obtain A(1)u +A(2)u =

u+ε

0

K(u+ε, r)ξ(r)dr− u

0

K(u, r)ξ(r)dr, which completes the proof of (3.7). As a consequence, from (3.5)–(3.7), we obtain

Sε(ξ )= T

0

1 ε

u+ε u

K+ξ(s)ds k

du.

On the other hand, using (3.4) and the definition of the maximal functionD, we get sup

0<εε0

1 ε

u+ε

u

K+ξ(s)ds

kMk|ξ|k17/12 sup

0<εε0

1 ε

u+ε

u

C(s)ds k

(3.14)

=Mk|ξ|k17/12Dk(u).

Therefore, using hypothesis(Hk3)and the dominated convergence theorem, we have

εlim0Sε(ξ )= T

0

(K+ξ(s))kds. (3.15)

Moreover, since|Sε(ξ )| ≤Mk|ξ|k17/12T

0 Dk(u)dufor all 0< εε0(see (3.14)), conditions (1) and (2) in Proposition4are fulfilled. Consequently,εkT

0 (Bu+εBu)kduconverges in(S) asε→0.

(13)

To complete the proof, it suffices to observe that the right-hand side of (3.15) is, by definition (see Lemma3), theS-transform ofT

0 B˙skds.

In [14], it is proved that under some additional hypotheses, the mappingtBt is differen- tiable from(0,)to(S)and that its derivative, denoted by B˙t, is a Hida distribution whose S-transform isK+ξ(t ).

3.2. Bidimensional case

LetW=(Wt)t∈Rbe a two-sided Brownian motion defined in the white noise probability space (S(R),B,P). We can consider two independent standard Brownian motions as follows: for t≥0, we setWt(1)=Wt andWt(2)=Wt.

In this section, we consider a bidimensional processB=(Bt(1), Bt(2))t0, whereB(1)andB(2) are independent Volterra processes of the form

Bt(i)= t

0

K(t, s)dWs(i), t≥0, i=1,2. (3.16) For simplicity only, we work with the same kernelKfor the two components.

First, using exactly the same lines of reasoning as in the proof of Lemma3, we get the follow- ing result.

Lemma 5. Let B=(Bt(1), Bt(2))t0 be given as above,with a kernel K satisfying the condi- tions(H1)and (H2).Assume,moreover,that C defined by(3.3)belongs toL2([0, T])for any T >0.We then have the following results:

(1) the function ξT

0 (u

0 K+ξ(y)dy)K+ξ(u)du is the S-transform of an element of (S),denoted byT

0 Bu(1) ˙Bu(2)du;

(2) the functionξT

0 K+ξ(u)K+ξ(u)duis theS-transform of an element of(S),de- noted byT

0 B˙u(1) ˙Bu(2)du.

We can now state the following result.

Proposition 6. LetB=(Bt(1), Bt(2))t0be given as above,with a kernelKsatisfying the condi- tions(H1),(H2)and(H23).The following convergences then hold:

T

0

Bu(1)Bu(2)+εBu(2) ε du−→(S)

ε0

T

0

Bu(1) ˙Bu(2)du, T

0

u 0

Bv(1)+εBv(1)

ε dv

Bu(2)+εBu(2) ε du−→(S)

ε0

T 0

Bu(1) ˙Bu(2)du, T

0

Bu(1)+εBu(1)

ε ×Bu(2)+εBu(2) ε du−→(S)

ε0

T

0

B˙u(1) ˙Bu(2)du.

(14)

Proof. Set

Gε= T

0

Bu(1)Bu(2)+εBu(2)

ε du=

T 0

Bu(1)Bu(2)+εBu(2)

ε du.

From linearity and property (2.3) of theS-transform, we have S( Gε)(ξ )=1

ε T

0

S Bu(1)

(ξ )S

Bu(2)+εBu(2) (ξ )du so that

S(Gε)(ξ )= T

0

u 0

K+ξ(y)dy 1

ε u+ε

u

K+ξ(x)dx

du.

Therefore, using (3.4) and (3.14), we can write

|S(Gε)(ξ )| ≤M2|ξ|217/12

T

0

u 0

C(t )dt

D(u)du

M2|ξ|217/12

T 0

u 0

D(t)dt

D(u)du

=1

2M2|ξ|217/12

T 0

D(u)du 2

T

2M2|ξ|217/12

T

0

D2(u)du.

Hence, by the dominated convergence theorem, we get

εlim0

S( Gε)(ξ )= T

0

u 0

K+ξ(y)dy

K+ξ(u)du. (3.17)

The right-hand side of (3.17) is theS-transform ofT

0 Bu(1) ˙Bu(2)du, due to Lemma5. Therefore, by Theorem2, we obtain the desired result in point (1).

The proofs of the other two convergences follow exactly the same lines of reasoning and are

therefore left to the reader.

4. Fractional Brownian motion case

4.1. One-dimensional case

Consider a (one-dimensional) fractional Brownian motion (fBm)B=(Bt)t0 of Hurst index H(0,1). This means thatBis a zero mean Gaussian process with covariance function

RH(t, s)=E(BtBs)=12(t2H+s2H− |ts|2H).

Références

Documents relatifs

This justifies our claim that the ∞-Mat´ern model is the extension of Mat´ern type III model to the case where φ may be countably infinite and its associated conflict graph has

To illustrate a possible use of our results, we study in Section 6 an exten- sion of the classical fractional Ornstein-Uhlenbeck process (see, e.g., Cherid- ito et al [7]) to the

Abstract We obtain a local limit theorem for the laws of a class of Brownian additive func- tionals and we apply this result to a penalisation problem... This is the content of

The identification of modal parameters damping ratio (ζ) and natural frequency (ω n ) uses the underlying linear dynamics of the beam, obtained when input has low level of

First observe that the proof of (IV-4.30) carries over to the present situation (the process X in (IV-4.30) is assumed to be special standard) and so if C e A(M) has a finite

chains., Ann. Nagaev, Some limit theorems for stationary Markov chains, Teor. Veretennikov, On the Poisson equation and diffusion approximation. Rebolledo, Central limit theorem

We prove functional central and non-central limit theorems for generalized varia- tions of the anisotropic d-parameter fractional Brownian sheet (fBs) for any natural number d..

It has not been possible to translate all results obtained by Chen to our case since in contrary to the discrete time case, the ξ n are no longer independent (in the case of